Sie sind auf Seite 1von 11

Journal of Non-Crystalline Solids 307310 (2002) 470480

www.elsevier.com/locate/jnoncrysol

Volume recovery of polystyrene: evolution of the


characteristic relaxation time
P. Bernazzani, S.L. Simon *

Department of Chemical Engineering, Texas Tech University, Box 43121, Lubbock, TX 79409-3121, USA

Abstract

We have developed a new experimental technique that uses intermittent temperature perturbations during volume
recovery in order to obtain quantitative information concerning the evolution of the characteristic relaxation time for
volume during structural recovery. The experiments are analogous to the intermittent creep experiments developed by
Struik. Using an automated capillary dilatometer and a polystyrene sample, the time-temperature history dependence
of the characteristic relaxation time for volume was investigated. Our preliminary results show that for the set of
temperature down jumps and memory experiments investigated, the characteristic relaxation time appears to depend
only on the instantaneous state of the material. However, the results are not totally consistent with the ToolNaray-
anaswamy model/KovacsAklonisHutchinsonRamos model.
2002 Elsevier Science B.V. All rights reserved.

1. Introduction perfect quench from an initial temperature T0


above Tg to an aging temperature, Ta , below Tg , the
Polymer glasses have been the object of a great ctive temperature is given by
deal of interest in the last 50 years [115]. Math-   R t b 
ematical models of structural recovery have been Tf T0 DT 1  e
 dt
0 s ; 1
developed, including Moynihans formulation [12]
of the ToolNarayanaswamy model (TNM) [13,14]
and the KovacsAklonisHutchinsonRamos where Tf is dened as the temperature at which
model (KAHR) [15], the latter of which is essen- a material would reach equilibrium if heated or
tially equivalent to the TNM model. These math- cooled along the glass line, DT Ta  T0 , t is the
ematical models are able to describe the time elapsed following the jump, s is the charac-
manifestations of glassy behavior [6,8,1214]. The teristic relaxation time describing the recovery and
TNM model uses the ctive temperature Tf to b is a parameter with values between 0 and 1 that
describe the changing structure of the glass. For a accounts for the non-exponentiality of the recov-
ery. The non-linearity of the recovery process is
accounted for by allowing s to be a function of
*
Corresponding author. Tel.: +1-806 742 3553; fax: +1-806
structure (Tf ), as in the following equation devel-
742 3552. oped by Narayanaswamy [14] and modied by
E-mail address: sindee.simon@coe.ttu.edu (S.L. Simon). Moynihan [12]:

0022-3093/02/$ - see front matter 2002 Elsevier Science B.V. All rights reserved.
PII: S 0 0 2 2 - 3 0 9 3 ( 0 2 ) 0 1 4 6 3 - 1
P. Bernazzani, S.L. Simon / Journal of Non-Crystalline Solids 307310 (2002) 470480 471

xDh 1  xDh He separately measured the creep response and


lns lnA ; 2 volume recovery during physical aging and then
RT RTf
cross-plotted the creep shift factor versus volume.
where lnA is a constant, R is the gas constant, x He concluded that the creep shift factor depended
partitions the dependence of s between tempera- only on the instantaneous volume of the sample.
ture and structure (Tf ) and Dh is the apparent ac- However due to scatter in the data, which may
tivation energy. We note that although s is referred have arisen in part from measuring the properties
to as a relaxation time, it is actually a retardation separately, his conclusions are arguable. Further-
time. more, the fact that the evolution of the charac-
In contrast to the TNM model which uses the teristic relaxation time for volume may not be the
ctive temperature as a measure of structure, the same as that for the mechanical relaxation due to
KAHR model describes structure in terms of a the possibility that there may be dierent time
departure from equilibrium (d): scales for dierent properties [2844] suggests the
V  V1 need to directly measure the evolution of the
d ; 3 characteristic relaxation time for volume.
V1
At this point it would be benecial to review the
where V is the instantaneous specic volume and experimental methodology used by Struik [1] to
V1 is the equilibrium specic volume. The depar- measure the instantaneous creep response during
ture from equilibrium (d) can easily be converted physical aging using sequential stress perturbation
to Tf , and vice versa, by the following equation: experiments. A schematic of Struiks experiment is
d shown in Fig. 1. After a down jump from T0 above
Tf Ta ; 4
Da Tg to Ta below Tg , intermittent stress perturbations
where Da is the change in thermal expansion co- are applied while the resulting strain is measured.
ecient from the equilibrium to the glassy states. The intermittent stress perturbations are 10 times
Using the relationship between Tf and d in Eq. (4), shorter than the aging time, te , so that the material
which assumes a perfect quench (otherwise To does not signicantly age during the perturbation.
should be replaced by Tf0 in the equation), Eq. (1) The time between each perturbation must be such
can be rewritten in terms of d: that the recovery from the last perturbation does
" Z not overlap the next perturbation. The recovery
t b #
dt time is generally two or three times the aging time.
d d0 exp  ; 5 Each stress perturbation or creep experiment gives
0 s
a snapshot of the creep response at a particular
where d0 is the departure from equilibrium im- aging time.
mediately after an ideal quench. Analogous to Struiks methodology we have
Although both the TNM/KAHR model can developed a new experiment to directly study the
describe the phenomena associated with the ki- characteristic relaxation time for volume during
netics of the glass transition, they are unable to aging. It makes use of intermittent temperature
simultaneously describe experiments performed perturbations during volume recovery to obtain
over a wide range of thermal histories and their quantitative information concerning the evolution
predictive capability is thus limited [6,8,1625]. of the characteristic volume relaxation time during
Simple corrections to the model provide little im- structural recovery. Fig. 2 shows the methodology
provement [8,19,23,25,26]. developed. Intermittent temperature perturbations
A basic assumption common to the TNM and are performed after an initial temperature jump
KAHR models is that the characteristic relaxation from T0 above Tg to an aging temperature, Ta ,
time, s, depends on the instantaneous state of the below Tg . Each temperature perturbation yields a
material, and not on its time-temperature history. snapshot of the volume response at a particular
Struik tested this assumption for polycarbonate, aging time. As a consequence, the characteristic
polyvinyl chloride and for polystyrene (PS) [27]. relaxation time for volume, s, can be obtained as a
472 P. Bernazzani, S.L. Simon / Journal of Non-Crystalline Solids 307310 (2002) 470480

Fig. 1. Experimental methodology used by Struik to measure the evolution of the mechanical characteristic relaxation time during
aging.

Fig. 2. Experimental methodology used in this paper to measure the evolution of the characteristic relaxation time for volume during
aging.

function of aging time. Again, the sequential per- relaxation time for volume recovery during the
turbations are short compared to the aging time, aging experiment.
such that the material does not age signicantly The paper is organized as follows. First, we
during the perturbations. Hence, the method al- detail the experimental methodology used in the
lows the direct determination of the characteristic work, including the thermal histories performed
P. Bernazzani, S.L. Simon / Journal of Non-Crystalline Solids 307310 (2002) 470480 473

and the method of data analysis. We include the brating the sample and dilatometer in a silicon oil
results of TNM/KAHR model calculations, which bath maintained at a temperature above Tg and
show that the methodology and data analysis are then quickly transferring the dilatometer to a bath
consistent with the model. We follow with pre- maintained at the aging temperature. For the tem-
liminary results, discussion, and conclusions. perature perturbations of 1 C, the dilatometer was
left in the bath and a 1 C step change was pro-
grammed. The specic thermal histories used are
2. Methodology detailed below. The standard deviation of the mean
temperature over a period of 3.5 days is 0.003 C;
2.1. Material however, the bath temperature is only reported to
0.01 C due to the data acquisition system used.
The material studied is a PS, Dylene 8 from The time for temperature equilibration was
Arco Polymers. The number-average molecular found to be 120 s for both temperature jumps
weight is 92 800 g/mol; the weight-average molec- (from one bath to another and for the temperature
ular weight is 221 000 g/mol; the z-average mo- perturbations). Time zero for the jumps is taken as
lecular weight is 423 000 g/mol [32]. It was the time at which the temperature change is initi-
obtained in pellet form. ated. Although some authors have advocated the
A dilatometric sample was obtained by sinter- use of various other time zeros [2,33,48], modeling
ing the polystyrene pellets in a 0.5-in. diameter work from our laboratory has shown that for the
tubular mold at 120 C under vacuum to minimize cooling history in the dilatometric experiments,
voids. A 0.125-in. hole was then drilled through taking time zero at the beginning of the nite
the center to allow a better contact with the mer- quench results in insignicant error in data com-
cury in the dilatometer. The nal mass of the pared to the results of a perfect quench for times
sample was 5 g. greater than approximately two times the thermal
equilibration time [49]. Taking time zero at the
2.2. Dilatometric studies beginning of the quench has been previously sug-
gested by Struik [50].
A capillary dilatometer, based on the design of
Plazek, was used [45]. The dilatometer consists of a 2.3. Thermal histories
stainless steel bulb containing the sample and
mercury as the surrounding uid. A precision The material was equilibrated at a temperature
bored Pyrex capillary tube is axed to the bulb. In T0 above Tg to erase all previous thermal histories.
order to automate the system, an aluminum oat For down jump experiments, the sample is quen-
is placed on top of the mercury and attached to a ched from T0 to 96 C and the evolution of volume
linear variable dierential transformer [46]. A at 96 C is monitored for 80 000 s. The down
computer records the output. The data acquisi- jump is shown in the schematic Fig. 3 by path ab.
tion is controlled using Labviewe software from For the memory experiments [2], the sample is
National Instruments. To eliminate recurring vi- quenched from T0 to a temperature T1 below Tg
brations a ltering technique was used: a Fast and kept at T1 for time t1 such that upon reheating
Fourier Transform of the data was performed to 96 C, the material is at approximately zero
every second over 200 data points and the higher departure from equilibrium. The evolution of the
frequencies were eliminated. The inverse transform volume at 96 C is then monitored for 80 000 s.
was then performed with minimal lag variation The path acde in the schematic Fig. 3 repre-
[47]. The result is analogous to a low pass lter. sents a memory experiment. It is noted that Fig. 3
The resolution of the system is better than 0:4  shows the down jump and memory experiments
105 cm3 /g. for a perfect quench. In reality the glass line after
Temperature jumps from equilibrium to the ag- a quench is lower than that indicated on the
ing temperature were accomplished by rst equili- schematic. However, as mentioned above, the
474 P. Bernazzani, S.L. Simon / Journal of Non-Crystalline Solids 307310 (2002) 470480

  R t b    R t b #
 dt  dt
Tf T0 DTa 1  e 0 s DTp 1  e tp s
;

6
where DTp Tp  Ta and tp is the time at which the
temperature perturbation is applied. Although se-
quential perturbations were performed, the time
between each one is at least twice the aging time so
that they do not aect one another and can be an-
alyzed independently. The time integral from 0 to t
of 1=s in Eq. (6) can be divided into two integrals:
Z t Z tp Z t
dt dt dt
7
0 s 0 s tp s

and the rst integral on the right side of the Eq. (7)
can be rewritten as a function of departure from
equilibrium dp , using Eq. (1):
Fig. 3. Schematic curve of the evolution of volume with tem- Z tp   1=b
dt d0
perature showing the thermal histories presented in this paper. ln ; 8
0 s dp
Table 1
Thermal histories for all experiments
where dp is the departure from equilibrium at Ta at
time tp when the perturbation experiment is star-
T0 (C) T1 (C) t1 (h)
ted. The second integral on the right side of Eq. (7)
Down jumps
can be simplied since s is essentially constant
104
102 during the perturbation due to the short duration
100 of the perturbation relative to the time of aging;
98 that is, the structural recovery taking place during
Memory experiments the perturbation does not cause a signicant shift
104 85 30 in the characteristic relaxation time. With this
104 80 70 assumption it follows that
104 75 160 Z t
dt t  tp
; 9
tp s sp
evolution of volume after a nite quench is indis-
tinguishable to that of a perfect quench for times where sp is the value of the relaxation time during
greater than twice the thermal equilibration time the perturbation. Substituting the expressions in
[49]. Table 1 shows the specic values of T0 , T1 and Eqs. (7)(9) into Eq. (6), the evolution of Tf during
t1 used in our experiments. Perturbations were the perturbation is then given by
2 h  i1=b
b 3
performed as in schematic Fig. 2 from 96 to 95 C.
d0 ttp
6  ln dp sp 7
2.4. Data analysis Tf T0 DTa 641  e
7
5

Eq. (1) shows the expected response to a single  tt b 


temperature jump from a temperature T0 to an  spp
DTp 1  e : 10
aging temperature Ta , in terms of ctive tempera-
ture Tf . In addition to the down jump, the tem- Eq. (10) can be rewritten in terms of departure
perature perturbation is taken into account by from equilibrium using Eq. (4). After simplica-
adding another term: tion, the departure from equilibrium during the
P. Bernazzani, S.L. Simon / Journal of Non-Crystalline Solids 307310 (2002) 470480 475

perturbation at the perturbation temperature can We note that the results of the analyses of the
be written perturbations performed at the shortest aging
2 h  i1=b
b 3 times (te  1500 and 3000 s) are not presented
6 
d0
ln d
ttp
sp  ttp
 b
7 because the time scale of the perturbation (being
dTp Da6 DTp e sp 7
p 
4DTa e 5: 0.1te ) was not long enough to give unique values of
sp and b.
11
2.5. Methodology validation
To derive the equation used to analyze the
perturbations from the memory experiments, an- To validate our data analysis (Eq. 11) we sim-
other term was added to Eq. (6) taking into ac- ulated data for down jump experiments using the
count the up-jump. TNM model and various model parameters. The
2  R b 3 simulations consisted of a down jump from 104 to

t
dt 96 C followed by three temperature perturbations
4
Tf T0 DTa 1  e 0 s 5
to 95 C at times 3600, 10 800 and 32 400 s from
2 R b 3 2 R b 3 the start of the initial jump. Eq. (11) was then used

t
dt 
t
dt to t the response to these perturbations in a
DT1 41  e t1 s 5 DTp 41  e tp sp 5; manner similar to that used for data obtained with
the automated dilatometer.
12 Fig. 4 shows the correlation between the char-
acteristic relaxation time values found using the
with t1 being the time at the beginning of the up-
methodology derived above with those expected
jump and DTa Ta  T0 ; DT1 T1  Ta ; DTp
for three dierent model simulations. The expected
Tp  T1 , using the same notation as in Fig. 2.
values are the average lnsp during the perturba-
Carrying on a development similar to the one
tions. The TNM parameters used in the modeling
above, we obtain
are shown on the gure as well as the line where
 
b b lnsp calculated lnsp expected. The error in sp
dTp Da DT0 e ABC DT1 eBC DTp eC ;
b

is at most 4% indicating that the data analysis is


13 consistent with the TNM model.
where
  1=b Z tp
d0 dt t  tp
A ln ; B ; C :
dp t1 s sp
The ts were performed using KaleidaGraph
from Synergy Software, which utilizes the Leven-
bergMarquardt tting algorithm. The values of
d0 =d were obtained from the data at the beginning
of each perturbation. The data were tted a rst
time allowing the non-exponentiality parameter, b,
to vary and tting for sp for the down jump ex-
periments. For the memory experiments, both sp
and B were obtained from the t. b was found
to vary from 0:48  0:03 to 0:55  0:05 to in a
non-systematic way, while B varies for each per-
turbation since it is a function of the time of the Fig. 4. Correlation between the characteristic relaxation time
perturbation. The b values were averaged (b values found using Eq. 11 for perturbations simulated during
0:52) and the data was retted using this value of b. volume recovery after jumping from 104 to 96 C.
476 P. Bernazzani, S.L. Simon / Journal of Non-Crystalline Solids 307310 (2002) 470480

3. Results and discussion

3.1. Cooling studies

Fig. 5 shows the specic volume versus tem-


perature curve obtained on cooling at a rate of
about 0.2 C/min. The thermal expansion coe-
cient is calculated to be 5:2  104 K1 in the liquid
state and 2:0  104 K1 in the glassy region. These
values agree with our previous work [32] and with
values [38,5154] found in the literature. The glass
transition temperature is determined as the inter-
cept of the glass and liquid lines. A value of 94.5 C
was found (at 0.2 C/min) which is similar to the
value of 94.8 C observed previously [32].
Fig. 6. Departure from equilibrium at 96 C during volume
3.2. Eect of temperature perturbation on underly- recovery of PS. The line is the volume recovery without tem-
perature perturbations, and the points show the response when
ing recovery sequential temperature perturbations were performed to 95 C
during the recovery. For both runs, T0 104 C and Ta 96
In order to use the perturbations as a probe of C.
the volume relaxation response, the underlying
volume relaxation should be unaected by the
temperature perturbations. Fig. 6 shows the vol- ure demonstrates that the perturbations do not
ume recovery of the PS sample for two runs. The aect the underlying recovery since the two runs
initial temperature, T0 , is 104 C and the aging are identical within experimental error. Note that
temperature is 96 C for both runs. The line rep- the responses to the perturbation up jumps (from
resents the response to a simple down jump while 95 back to 96 C) were not recorded.
the data points are the response to a down jump To further conrm that the temperature per-
with temperature perturbations to 95 C. The g- turbations do not aect the underlying volume re-
covery, we t the underlying response with the
TNM model and compared the results to that ob-
tained in previous work [32] on the same polymer.
The results are summarized in Table 2. Both the
present and previous ts were performed assuming
a nite (rather than innite) temperature quench.
The values obtained for the tting parameters
Dh=R, x and b are similar to those obtained in
Ref. [32], indicating that the perturbations do not

Table 2
TNM parameters comparison
Parameter Model ts of the Results from
underlying recovery Ref. [32]
Dh=R (kK) 138.3 146.1
x 0.2 0.1
b 0.80 0.70
lnA=s 366.9 387.40
Fig. 5. Volume evolution of polystyrene on cooling at 0.2 C/
logs1 =s at 96 C 3.4 3.6
min.
P. Bernazzani, S.L. Simon / Journal of Non-Crystalline Solids 307310 (2002) 470480 477

signicantly aect the underlying volume recovery


of our sample. The value of lnA diers slightly
due to the dierence in Dh=R. The value of the
logarithm of the equilibrium relaxation time cal-
culated at 96 C is lower by 0.2 in this work.

3.3. Volume recovery

The volume recovery response depends on the


magnitude of the temperature jump as indicated
by Eq. (1). Fig. 7 shows the recovery of the sample
at 96 C for initial temperatures (T0 ) ranging from
104 to 98 C. Temperature perturbations of 1 C
from 96 C to 95 C were performed during each
run. The departure from equilibrium is greatest for
the highest T0 . The curves approach equilibrium Fig. 8. Departure from equilibrium at 96 C for polystyrene in
response to memory experiments with the temperature history
asymptotically and overlap at long times as ex-
104T1 96 C and to a down-jump from 10496 C. Tempera-
pected, based on the extensive data set of Kovacs ture perturbation were performed to 95 C.
[2]. Before analyzing the response to the tempera-
ture perturbations, we rst look at the results of
the memory experiments. ture from equilibrium at 96 C is initially near
For memory experiments, the volumetric re- zero, then increases and nally returns to zero. The
sponse is more complicated, as shown in Fig. 8. overshoot increases as T1 decreases. The responses
The thermal histories for the memory experiments converge to that of the down jump and approach
are indicated on the gure. Also shown is the re- equilibrium asymptotically at long times. These
sponse to the simple down jump to 96 C with results are consistent with those of Kovacs who
T0 104 C. In all cases, the volume response was attributed the overshoot to a memory eect im-
measured at 96 and 1 C perturbations were per- plying the existence of a distribution of charac-
formed. For the memory experiment, the depar- teristic relaxation times [2].

3.4. Response to perturbations

We are interested in whether the characteristic


relaxation time depends only on the instantaneous
state of the material or on the time/temperature
history. To get this information, we examine the
volume response to the temperature perturbation.
Fig. 9 shows the departure from equilibrium d
versus time of the perturbation for a down jump
from 104 to 96 C and for a memory experiment
from 1048596 C for several perturbations. d is
dened here, as the departure from equilibrium at
95 C. We emphasize that in order to examine only
the response to the perturbation, we subtracted the
underlying relaxation from the response. Fig. 9
Fig. 7. Departure from equilibrium at 96 C during volume
shows that, as expected, the volumetric response to
recovery of PS for down jumps from T0 to Ta 96 C. Tem- the 1 C temperature perturbation changes with d,
perature perturbations were performed to 95 C. shifting to longer times with decreasing d. When
478 P. Bernazzani, S.L. Simon / Journal of Non-Crystalline Solids 307310 (2002) 470480

tting the two experiments, along with the tting


error, are also shown.
By combining all of the data sets, we can now
examine the dependence of the characteristic re-
laxation time for volume on the value of d aver-
aged over the duration of the perturbation for
various thermal histories, as shown in Fig. 10. The
error bars in the gure represent the error in the t
of sp obtained using Eq. (11) or Eq. (13) for b
0:52. Independent of the thermal history, the data
suggest that the characteristic relaxation time does
depend on the departure from equilibrium d, as the
TNM and KAHR models assume.
It is noted that in the initial analysis, the data
Fig. 9. Departure from equilibrium at the perturbation tem-
points with the history 1048096 C were nearly a
perature (95 C) versus time of the perturbation for a down full logarithmic unit lower than the rest of the
jump from 104 to 96 C and for a memory experiment from data points and the d versus log time perturba-
1048596 C for several perturbations performed during aging tion responses were systematically shifted to values
at 96 C. For clarity the underlying recoveries have been sub- smaller than expected. Shifting these up by 6 
tracted. The continuous and dashed lines represent the ts to
the KWW equation as explained in the text, for the down jump
105 in d resulted in perturbation response curves
and memory experiment respectively. and s values that agreed with the rest of the data.
The systematic error cannot be explained by an
error in the temperature measurements. Using the
the system achieves equilibrium, the shape of the value of the thermal expansivity in the glass ob-
response and therefore the value of sp are identical tained from previous work [32] we calculate that
for all curves independent of the thermal history of to obtain a shift of 6  105 in d would require a
the sample. More importantly, the volumetric re-
sponse to the perturbations appears to be inde-
pendent of the path, i.e., the down jump and
memory experiments show the same response to
the perturbations at a given initial departure from
equilibrium. The data are representative of the
temperature histories examined in this work.
We can quantify the evolution of the volumetric
response to the 1 C perturbation by tting the
data as described earlier. However, in this exam-
ple, since the underlying recovery curve have been
subtracted from the perturbation responses, Eqs.
(11) and (13) simplify to a simple Kolrausch
WilliamsWatts equation (KWW) [55,56] for the
case where the perturbation does not aect the
underlying response:
 b
ttp
 Fig. 10. Dependence of the characteristic relaxation time for
dp DTp Da e : 14
sp
volume on the departure from equilibrium for aging at 96 C
following various thermal histories. The solid line represents the
The ts to the KWW are shown as lines for the linear t of the data. The dashed line represents the result based
down jump and as dashed lines for the memory on the ts of the underlying curves with the parameters shown
experiment in Fig. 9. The sp values averaged from in Table 2.
P. Bernazzani, S.L. Simon / Journal of Non-Crystalline Solids 307310 (2002) 470480 479

change in the perturbation temperature of 0.3 C, models, although it is by no means necessary, that
which was not observed. Although it is unclear the relationship between lns and Tf or between
why we had this problem with one of the seven lns and d is linear. We note that Schick and co-
temperature histories we examined, we believe the workers [60] have recently found that they were
correction made is reasonable. Backing this up is only able to well describe the loss and storage heat
the fact that the data from the memory experi- capacity obtained from slow temperature-modu-
ments in which T1 was lower (75 C) and higher (85 lated DSC cooling scans by using a nonlinear re-
C) both lie on the lns versus d line where the lationship based on Adam and Gibbs.
majority of data points are located.
In Fig. 10, the slope of the lns versus d rela-
tionship is related to the KAHR parameters: 4. Conclusion
 
d lns 1  xH This paper presents the development of a new
 ; 15 methodology capable for the rst time of fol-
dd Ta Da
lowing the evolution of the characteristic relax-
where H Dh=RT 2 and T is the temperature. In ation time for volume during the volume recovery.
Fig. 10, this slope is 7920  1000. A value of Preliminary results are reported for a PS glass.
2400  800 (the dashed line in Fig. 10) was cal- This methodology, analogous to the protocol
culated using our TNM model ts to the under- popularized by Struik for creep measurements,
lying recoveries and using Da 3:2  102 K1 uses sequential temperature perturbations that are
(Table 2). Alternatively, using the value of Dh=R of performed during the structural recovery of the
138.3 kK, we calculated x 1:6 from the slope of sample. The perturbations do not aect the un-
lns versus d. The values obtained from this derlying volume recovery. For the thermal treat-
analysis and the t to the underlying recovery are ments investigated, the characteristic relaxation
inconsistent. A possible explanation could be that time for volume appears to depend on the instan-
the use of temperature perturbations to follow the taneous state of the material. The slope of  lns
characteristic relaxation time focuses on the short versus d is 7920, considerably larger than expected
time scale response. As a consequence, the TNM based on this and previous modeling of intrinsic
parameters in Table 2, which were found by ana- isotherms of the same material. In addition, the
lyzing the complete relaxation curve, might be analysis suggests that the relationship between
underestimating the rate of relaxation at very short lns and d may not be linear as d varies from d0 to
times. This has been reported for systems far from 0 in contrast to the assumption commonly made in
equilibrium where the TNM model is incapable of the TNM/KAHR models.
tting the data [19,57].
Another explanation for the apparent discrep- Acknowledgements
ancy between the strong dependence of lns on d
and the value of Dh=R for the TNM equation is that The authors gratefully acknowledge fruitful
the relationship between the characteristic relax- discussions with Cornelius Moynihan, Mark Edi-
ation time and the departure from equilibrium is ger and Gregory McKenna concerning this work.
non-linear. Such non-linear relationships have Partial nancial support from the American Chemi-
been derived from AdamGibbs theory [58] by cal Society Petroleum Research Fund 36003-AC7
both Hodge and Scherer [8,59]. In other words, a is also gratefully acknowledged.
linear extrapolation of the lns versus d relation-
ship leads to values of lns at d0 which are much
smaller than would be expected. A natural way to References
satisfy this is with a relationship between the two
variables that is not linear. This contrasts with the [1] L.C.E. Struik, Physical Aging in Amorphous Polymers and
assumption commonly made in the TNM/KAHR Other Materials, Elsevier Scientic, New York, 1978.
480 P. Bernazzani, S.L. Simon / Journal of Non-Crystalline Solids 307310 (2002) 470480

[2] A. Kovacs, J. Adv. Polym. Sci. 3 (1963) 394. [29] J.M.G. Cowie, R. Ferguson, S. Harris, I.J. McEwen,
[3] J.M.G. Cowie, J. Macromol. Sci. Phys. B 18 (1980) 569. Polymer 39 (1998) 4393.
[4] D.J. Plazek, G.C. Berry, Glass Sci. Technol. 3 (1986) 363. [30] J.M.G. Cowie, S. Harris, I.J. McEwen, Macromolecules 31
[5] J. OReilly, CRC Crit. Rec. Solid Mater. Sci. 13 (1987) 259. (1998) 2611.
[6] G.B. McKenna, in: C. Booth, C. Price (Eds.), Compre- [31] J.M.G. Cowie, S. Elliott, R. Ferguson, R. Simha, Polym.
hensive Polymer Science, Polymer Properties, vol. 12, Commun. 28 (1987) 298.
Pergamon, Oxford, 1989. [32] S.L. Simon, J.W. Sobieski, D.J. Plazek, Polymer 42 (2001)
[7] G.W. Scherer, J. Non-Cryst. Solids 123 (1990) 75. 2555.
[8] I.M. Hodge, J. Non-Cryst. Solids 169 (1994) 211. [33] S.L. Simon, D.J. Plazek, J.W. Sobieski, E.T. McGregor, J.
[9] J. Mijovic, J. Polym. Eng. Sci. 34 (1994) 381. Polym. Sci. Part B 35 (1997) 929.
[10] J.M. Hutchinson, Prog. Polym. Sci. 20 (1995) 703. [34] G.B. McKenna, Y. Leterrier, C.R. Schultheisz, Polym.
[11] J.A. Forrest, R.A.L. Jones, in: A. Karim, S. Kumar (Eds.), Eng. Sci. 35 (1995) 403.
Polymer Surfaces, Interfaces and Thin Films, World [35] C.R. Schultheisz, G.B. McKenna, Y. Leterrier, E. Stefanis,
Scientic, Singapore, 2000. in: Conference Papers Society for Experimental Mechanics,
[12] C.T. Moynihan, P.B. Macedo, C.J. Montrose, P.K. Gupta, 1114 June 1995, Grand Rapid, MI.
M.A. DeBolt, J.F. Dill, B.E. Dom, P.W. Drake, A.J. [36] G.W. Scherer, Relaxation in Glass and Composites, John
Easteal, P.B. Elterman, R.P. Moeller, H. Sasabe, J.A. Wiley, New York, 1986.
Wilder, N.Y. Ann. Acad. Sci. 15 (1976) 279. [37] I. Echeverria,P. Kolek, S.L. Simon, D.J. Plazek, J. Chem.
[13] A.Q. Tool, J. Am. Ceram. Soc. 29 (1946) 240. Phys., in preparation.
[14] O.S. Narayanaswamy, J. Am. Ceram. Soc. 54 (1971) 491. [38] R. Greiner, F.R. Schwarzl, Rheol. Acta 23 (1984) 378.
[15] A.J. Kovacs, J.M. Aklonis, J.M. Hutchinson, A.R. Ramos, [39] R.J. Roe, G.M. Millman, Polym. Eng. Sci. 23 (1983) 318.
J. Polym. Sci. Polym. Phys. Ed. 17 (1979) 1097. [40] H. Sasabe, C.T. Moynihan, J. Polym. Sci. B 16 (1978)
[16] C.T. Moynihan, S.N. Crichton, S.M. Opalka, J. Non- 1447.
Cryst. Solids 131133 (1991) 420. [41] S.E.B. Petrie, in: G. Allen, S.E.B. Petrie (Ed.), Physical
[17] C.T. Moynihan, Rev. Mineral 32 (1995) 1. Structure of the Amorphous State, Marcel Dekker, New
[18] G.B. McKenna, C.A. Angell, J. Non-Cryst. Solids 133 York, 1977.
(1991) 528. [42] J.M. Hutchinson, S. Smith, B. Horne, G.M. Gourlay,
[19] G.W. Scherer, J. Am. Ceram. Soc. 69 (1986) 374. Macromolecules 32 (1999) 5057.
[20] I.M. Hodge, G.S. Huvard, Macromolecules 16 (1983) [43] J. Perez, J.Y. Cavaille, Makromol. Chem. 192 (1991) 2141.
371. [44] K. Adachi, T. Kotaka, Polym. J. 14 (1982) 959.
[21] J.J. Tribone, J.M. OReilly, J.J. Greener, Polym. Sci. B 27 [45] C.A. Bero, D.J. Plazek, J. Polym. Sci. B 29 (1991) 39.
(1989) 837. [46] R.S. Duran, G.B. McKenna, J. Rheol. 34 (1990) 813.
[22] J.M. OReilly, I.M. Hodge, J. Non-Cryst. Solids 131133 [47] R.J. OHalloran, D.E. Smith, Anal. Chem. 50 (1978) 1391.
(1991) 451. [48] M.M. Santore, R.S. Duran, G.B. McKenna, Polymer 32
[23] S.L. Simon, Macromolecules 30 (1997) 4056. (1991) 2377.
[24] G. Medvedev, J.M. Caruthers, Presented at the Society of [49] J.W. Sobieski, PhD thesis, University of Pittsburgh,
Rheology 71st Annual Meeting, Madison, WI, 1721 December 1999.
October 1999; also presentation at APS, Seattle, WA, 12 [50] L.C.E. Struik, Polymer 28 (1997) 1869.
16 March 2001. [51] G. Braun, A. Kovacs, J. Phys. Chem. Glasses 4 (1963) 152.
[25] S. Rekhson, J.-P. Ducroux, in: L.D. Pye, W.C. La Course, [52] H.H.D. Lee, F.J. McGarry, Polymer 34 (1993) 4267.
H.J. Stevens (Eds.), The Physics of Non-Crystalline Solids, [53] K. Ueberreiter, G. Kanig, J. Colloid Sci. 7 (1953) 569.
Taylor and Francis, London, 1992, p. 315. [54] J.E. McKinney, R. Simha, J. Res. Nat. Bur. Stand. A 81A
[26] C.T. Moynihan, in: Oral Presentation at Int. Conf. on (1977) 283.
Relaxations in Complex Systems, Crete, June 1990, see also [55] F. Kolrausch, Prog. Ann. Phys. 12 (1847) 393.
Ref. [5] for details. [56] G. Williams, D.C. Watts, Trans. Farad. Soc. 66 (1970) 80.
[27] L.C.E. Struik, Polymer 29 (8) (1988) 1347. [57] J. Huang, P.K. Gupta, J. Non-Cryst. Solids 151 (1992) 175.
[28] G.B. McKenna, C.R. Schultheisz, Y. Leterrier, in: Con- [58] G. Adam, J.H. Gibbs, J. Chem. Phys. 43 (1) (1965) 139.
ference Papers, 9th International Conference on Deforma- [59] G.W. Scherer, J. Am. Ceram. Soc. 67 (1984) 504.
tion, Yield and Fracture of Polymers, Cambridge, UK, [60] S. Weyer, M. Merzlyakov, C. Schick, Thermochim. Acta
April 1994, p. 31/1. 377 (1&2) (2001) 85.

Das könnte Ihnen auch gefallen